首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
2.
3.
4.
During yeast sporulation, a forespore membrane (FSM) initiates at each spindle-pole body and extends to form the spore envelope. We used Schizosaccharomyces pombe to investigate the role of septins during this process. During the prior conjugation of haploid cells, the four vegetatively expressed septins (Spn1, Spn2, Spn3, and Spn4) coassemble at the fusion site and are necessary for its normal morphogenesis. Sporulation involves a different set of four septins (Spn2, Spn5, Spn6, and the atypical Spn7) that does not include the core subunits of the vegetative septin complex. The four sporulation septins form a complex in vitro and colocalize interdependently to a ring-shaped structure along each FSM, and septin mutations result in disoriented FSM extension. The septins and the leading-edge proteins appear to function in parallel to orient FSM extension. Spn2 and Spn7 bind to phosphatidylinositol 4-phosphate [PtdIns(4)P] in vitro, and PtdIns(4)P is enriched in the FSMs, suggesting that septins bind to the FSMs via this lipid. Cells expressing a mutant Spn2 protein unable to bind PtdIns(4)P still form extended septin structures, but these structures fail to associate with the FSMs, which are frequently disoriented. Thus, septins appear to form a scaffold that helps to guide the oriented extension of the FSM.Yeast sporulation is a developmental process that involves multiple, sequential events that need to be tightly coordinated (59, 68). In the fission yeast Schizosaccharomyces pombe, when cells of opposite mating type (h+ and h) are mixed and shifted to conditions of nitrogen starvation, cell fusion and karyogamy occur to form a diploid zygote, which then undergoes premeiotic DNA replication, the two meiotic divisions, formation of the spore envelopes (comprising the plasma membrane and a specialized cell wall), and maturation of the spores (74, 81). At the onset of meiosis II, precursors of the spore envelopes, the forespore membranes (FSMs), are formed by the fusion of vesicles at the cytoplasmic surface of each spindle-pole body (SPB) and then extend to engulf the four nuclear lobes (the nuclear envelope does not break down during meiosis), thus capturing the haploid nuclei, along with associated cytoplasm and organelles, to form the nascent spores (55, 68, 81). How the FSMs recognize and interact with the nuclear envelope, extend in a properly oriented manner, and close to form uniformly sized spherical spores is not understood, and study of this model system should also help to elucidate the more general question of how membranes obtain their shapes in vivo.It has been shown that both the SPB and the vesicle trafficking system play important roles in the formation and development of the FSM and of its counterpart in the budding yeast Saccharomyces cerevisiae, the prospore membrane (PSM). In S. pombe, the SPB changes its shape from a compact dot to a crescent at metaphase of meiosis II (26, 29), and its outer plaque acquires meiosis-specific components such as Spo2, Spo13, and Spo15 (30, 57, 68). This modified outer plaque is required for the initiation of FSM assembly. In S. cerevisiae, it is well established that various secretory (SEC) gene products are required for PSM formation (58, 59). Similarly, proteins presumably involved in the docking and/or fusion of post-Golgi vesicles and organelles in S. pombe, such as the syntaxin-1A Psy1, the SNAP-25 homologue Sec9, and the Rab7 GTPase homologue Ypt7, are also required for proper FSM extension (34, 53, 54). Consistent with this hypothesis, Psy1 disappears from the plasma membrane upon exit from meiosis I and reappears in the nascent FSM.Phosphoinositide-mediated membrane trafficking also contributes to the development of the FSM. Pik3/Vps34 is a phosphatidylinositol 3-kinase whose product is phosphatidylinositol 3-phosphate [PtdIns(3)P] (35, 72). S. pombe cells lacking this protein exhibit defects in various steps of FSM formation, such as aberrant starting positions for extension, disoriented extension and/or failure of closure, and the formation of spore-like bodies near, rather than surrounding, the nuclei, suggesting that Pik3 plays multiple roles during sporulation (61). The targets of PtdIns(3)P during sporulation appear to include two sorting nexins, Vps5 and Vps17, and the FYVE domain-containing protein Sst4/Vps27. vps5Δ and vps17Δ mutant cells share some of the phenotypes of pik3Δ cells (38). sst4Δ cells also share some of the phenotypes of pik3Δ cells but are distinct from vps5Δ and vps17Δ cells, consistent with the hypothesis that Pik3 has multiple roles during sporulation (62).Membrane trafficking processes alone do not seem sufficient to explain how the FSMs and PSMs extend around and engulf the nuclei, suggesting that some other mechanism(s) must regulate and orient FSM/PSM extension. The observation that the FSM is attached to the SPB until formation of the immature spore is complete (68) suggests that the SPB may regulate FSM extension. In addition, the leading edge of the S. cerevisiae PSM is coated with a complex of proteins (the LEPs) that appear to be involved in PSM extension (51, 59). S. pombe Meu14 also localizes to the leading edge of the FSM, and deletion of meu14 causes aberrant FSM formation in addition to a failure in SPB modification (60). However, it has remained unclear whether the SPB- and LEP-based mechanisms are sufficient to account for the formation of closed FSMs and PSMs of proper size and position (relative to the nuclear envelope), and evidence from S. cerevisiae has suggested that the septin proteins may also be involved.The septins are a conserved family of GTP-binding proteins that were first identified in S. cerevisiae by analysis of the cytokinesis-defective cdc3, cdc10, cdc11, and cdc12 mutants (41). Cdc3, Cdc10, Cdc11, and Cdc12 are related to each other in sequence and form an oligomeric complex that localizes to a ring in close apposition to the plasma membrane at the mother-bud neck in vegetative cells (12, 20, 25, 41, 47, 77). The septin ring appears to be filamentous in vivo (12), and indeed, the septins from both yeast (11, 20) and metazoans (31, 36, 69) can form filaments in vitro. The yeast septin ring appears to form a scaffold for the localization and organization of a wide variety of other proteins (8, 22), and it forms a diffusion barrier that constrains movement of membrane proteins through the neck region (7, 8, 73). In metazoan cells, the septins are involved in cytokinesis but are also implicated in a variety of other cellular processes, such as vesicular transport, organization of the actin and microtubule cytoskeletons, and oncogenesis (27, 70).In S. cerevisiae, a fifth septin (Shs1) is also expressed in vegetative cells, but the remaining two septin genes, SPR3 and SPR28, are expressed at detectable levels only during sporulation (15, 17). In addition, at least some of the vegetatively expressed septins are also present in sporulating cells (17, 48), and one of them (Cdc10) is expressed at much higher levels there than in vegetative cells (32). The septins present during sporulation are associated with the PSM (15, 17, 48, 51), and their normal organization there depends on the Gip1-Glc7 protein phosphatase complex (71). However, it has been difficult to gain insight into the precise roles of the septins during sporulation in S. cerevisiae (59), because some septins are essential for viability during vegetative growth, and the viable mutants have only mild phenotypes during sporulation (15, 17), possibly because of functional redundancy among the multiple septins.S. pombe seemed likely to provide a better opportunity for investigating the role of septins during spore formation. There are seven septin genes (spn1+ to spn7+) in this organism (23, 41, 63). Four of these genes (spn1+ to spn4+) are expressed in vegetative cells, and their products form a hetero-oligomeric complex that assembles during cytokinesis into a ring at the division site (2, 3, 10, 76, 79). The septin ring is important for proper targeting of endoglucanases to the division site (44), and septin mutants show a corresponding delay in cell separation (10, 41, 44, 76). However, even the spn1Δ spn2Δ spn3Δ spn4Δ quadruple mutant is viable and grows nearly as rapidly as the wild type (our unpublished results), a circumstance that greatly facilitates studies of the septins'' role during sporulation.spn5+, spn6+, and spn7+ are expressed at detectable levels only during sporulation (1, 45, 78; our unpublished results), and spn2+, like its orthologue CDC10 (see above), is strongly induced (45), but the roles of the S. pombe septins in sporulation have not previously been investigated. In this study, we show that the septins are important for the orientation of FSM extension, suggesting that the septins may have a more general role in dynamic membrane organization and shape determination.  相似文献   

5.
Deficiency in both ATM and the DNA-dependent protein kinase catalytic subunit (DNA-PKcs) is synthetically lethal in developing mouse embryos. Using mice that phenocopy diverse aspects of Atm deficiency, we have analyzed the genetic requirements for embryonic lethality in the absence of functional DNA-PKcs. Similar to the loss of ATM, hypomorphic mutations of Mre11 (Mre11ATLD1) led to synthetic lethality when juxtaposed with DNA-PKcs deficiency (Prkdcscid). In contrast, the more moderate DNA double-strand break response defects associated with the Nbs1ΔB allele permitted viability of some Nbs1ΔB/ΔB Prkdcscid/scid embryos. Cell cultures from Nbs1ΔB/ΔB Prkdcscid/scid embryos displayed severe defects, including premature senescence, mitotic aberrations, sensitivity to ionizing radiation, altered checkpoint responses, and increased chromosome instability. The known functions of DNA-PKcs in the regulation of Artemis nuclease activity or nonhomologous end joining-mediated repair do not appear to underlie the severe genetic interaction. Our results reveal a role for DNA-PKcs in the maintenance of S/G2-phase chromosome stability and in the induction of cell cycle checkpoint responses.The Mre11 complex, consisting of Mre11, Rad50, and Nbs1 (Xrs2 in Saccharomyces cerevisiae), is involved in diverse aspects of DNA double-strand break (DSB) metabolism. The Mre11 complex acts as a DSB sensor, mediates cell cycle checkpoint arrest and apoptosis, and promotes DSB repair (47, 48). The influence of the Mre11 complex on DSB responses is attributable partly to its influence on ataxia-telangiectasia mutated (ATM) kinase activity (29). ATM is a central signal transducer in the response to DSBs and is required for arrest throughout the cell cycle, as well as the efficient execution of apoptosis in response to many types of genotoxic stress (43).The Mre11 complex is required for ATM activation and governs the phosphorylation of ATM substrates such as SMC1, Chk2, and BID (4, 6, 26, 47, 49, 51). The C terminus of Nbs1 interacts with ATM and plays an important role in facilitating a subset of these events, particularly those important for apoptosis (11, 14, 47, 58). However, ATM makes multiple functional contacts with members of the Mre11 complex. Nbs1, Mre11, and Rad50 are all ATM substrates, and many aspects of ATM checkpoint signaling are impaired by hypomorphic Mre11 and Nbs1 mutations that do not affect the ATM binding domain in the C terminus of Nbs1 (32, 36, 52, 54).Several molecular and genetic observations support the view that the Mre11 complex''s role in preserving genome stability is particularly relevant to the S and G2 phases of the cell cycle (3, 56). The complex, predominantly nucleoplasmic in G1 cells, becomes predominantly chromatin associated and colocalizes with PCNA throughout S phase (35, 38). This association is a likely prerequisite for the complex''s influence on DNA damage signaling as well as DNA repair.Cell cultures established with samples from patients with Nijmegen breakage syndrome (NBS1 hypomorphism) and ataxia-telangiectasia-like disorder (MRE11 hypomorphism) exhibit checkpoint defects in S phase and at the G2/M transition, while the G1/S transition is relatively unaffected. These checkpoint defects are correlated with reduced Mre11 complex chromatin association both in human cells and in mouse models of Nijmegen breakage syndrome and ataxia-telangiectasia-like disorder (5, 45, 49, 52). Chromosomal aberrations arising in these cells are predominantly chromatid type breaks, consistent with impaired metabolism of DNA replication-associated DNA breaks (49, 52).Further supporting a predominant role for the Mre11 complex in S phase is the observation that its primary role in DSB repair is the promotion of recombination between sister chromatids (3, 24). Structural and genetic evidence that the Mre11 complex effects molecular bridging between DNA duplexes offers a mechanistic basis for this observation (10, 23, 53). Molecular bridging by the Mre11 complex may also contribute to its influence on nonhomologous end joining (NHEJ) (12, 34, 57). Collectively, these data strongly support the view that the Mre11 complex''s checkpoint and DSB repair functions are manifested predominantly in the S and G2 phases of the cell cycle.Although the Mre11 complex and ATM function in the same arm of the DNA damage response, ATM deficiency is lethal in hypomorphic Mre11 and Nbs1 mutants (Mre11ATLD1/ATLD1 and Nbs1ΔB/ΔB mice, respectively) (49, 52), suggesting that aspects of ATM function are Mre11 complex independent. ATM deficiency is also synthetically lethal with mutations in Prkdc, the gene encoding the catalytic subunit of the DNA-dependent protein kinase (DNA-PKcs) that is mutated in mice with severe combined immunodeficiency (Prkdcscid mice) (22, 42). DNA-PKcs is an ATM paralog required for NHEJ, which appears to be the predominant mode of DSB repair in G1 cells (16).Defective NHEJ is unlikely to be the basis for the embryonic lethality of Prkdc/ Atm/ or Prkdcscid/scid Atm/ mice, as loss of ATM rescues the late embryonic lethality of both DNA ligase IV (Lig4) and XRCC4 null embryos, which have more severe NHEJ defects than Prkdcscid mice abolished by the Atm/ genotype (31, 42). These observations argue that the DNA-PKcs functions required for viability in the absence of ATM do not include NHEJ.To address this issue, we crossed Mre11ATLD1/ATLD1 and Nbs1ΔB/ΔB mice with Prkdcscid/scid mice. As these Mre11 complex hypomorphs do not completely phenocopy ATM deficiency, we reasoned that double-mutant animals would be viable and thus provide a venue in which to examine the functional relationship between the Mre11 complex/ATM arm of the DNA damage response and DNA-PKcs. Whereas the Mre11ATLD1/ATLD1 mutation was synthetically lethal with the Prkdcscid/scid genotype, some Nbs1ΔB/ΔB Prkdcscid/scid mice were born, consistent with the more moderate DNA damage response defects associated with the Nbs1ΔB allele than with the Mre11ATLD1 allele (48). Nbs1ΔB/ΔB Prkdcscid/scid embryos were born at drastically reduced Mendelian ratios, displayed gross developmental defects, and were severely runted. Nbs1ΔB/ΔB Prkdcscid/scid cell cultures exhibited profound chromosome instability, growth defects, and increased sensitivity to ionizing radiation (IR). DNA repair defects associated with DNA-PKcs deficiency did not appear to underlie the observed phenotypic synergy. Rather, the data suggest a novel regulatory function of DNA-PKcs in the maintenance of chromosomal stability during the S and G2 phases of the cell cycle.  相似文献   

6.
7.
During the course of infection, transmitted HIV-1 isolates that initially use CCR5 can acquire the ability to use CXCR4, which is associated with an accelerated progression to AIDS. Although this coreceptor switch is often associated with mutations in the stem of the viral envelope (Env) V3 loop, domains outside V3 can also play a role, and the underlying mechanisms and structural basis for how X4 tropism is acquired remain unknown. In this study we used a V3 truncated R5-tropic Env as a starting point to derive two X4-tropic Envs, termed ΔV3-X4A.c5 and ΔV3-X4B.c7, which took distinct molecular pathways for this change. The ΔV3-X4A.c5 Env clone acquired a 7-amino-acid insertion in V3 that included three positively charged residues, reestablishing an interaction with the CXCR4 extracellular loops (ECLs) and rendering it highly susceptible to the CXCR4 antagonist AMD3100. In contrast, the ΔV3-X4B.c7 Env maintained the V3 truncation but acquired mutations outside V3 that were critical for X4 tropism. In contrast to ΔV3-X4A.c5, ΔV3-X4B.c7 showed increased dependence on the CXCR4 N terminus (NT) and was completely resistant to AMD3100. These results indicate that HIV-1 X4 coreceptor switching can involve (i) V3 loop mutations that establish interactions with the CXCR4 ECLs, and/or (ii) mutations outside V3 that enhance interactions with the CXCR4 NT. The cooperative contributions of CXCR4 NT and ECL interactions with gp120 in acquiring X4 tropism likely impart flexibility on pathways for viral evolution and suggest novel approaches to isolate these interactions for drug discovery.For human immunodeficiency virus type I (HIV-1) to enter a target cell, the gp120 subunit of the viral envelope glycoprotein (Env) must engage CD4 and a coreceptor on the cell surface. Although numerous coreceptors have been identified in vitro, the two most important coreceptors in vivo are the CCR5 (3, 11, 19, 22, 24) and CXCR4 (27) chemokine receptors. HIV-1 variants that can use only CCR5 (R5 viruses) are critical for HIV-1 transmission and predominate during the early stages of infection (86, 90). The importance of CCR5 for HIV-1 transmission is underscored by the fact that individuals bearing a homozygous 32-bp deletion in the CCR5 gene (ccr5-Δ32) are largely resistant to HIV-1 infection (15, 49, 84). Although R5 viruses typically persist into late disease stages, viruses that can use CXCR4, either alone (X4 viruses) or in addition to CCR5 (R5X4 viruses), emerge in approximately 50% of individuals infected with subtype B or D viruses (12, 39, 44). Although not required for disease progression, the appearance of X4 and/or R5X4 viruses is associated with a more rapid depletion of CD4+ cells in peripheral blood and faster progression to AIDS (12, 44, 77, 86). However, it remains unclear whether these viruses are a cause or a consequence of accelerated CD4+ T cell decline (57). The emergence of CXCR4-using viruses has also complicated the use of small-molecule CCR5 antagonists as anti-HIV-therapeutics as these compounds can select for the outgrowth of X4 or R5X4 escape variants (93).Following triggering by CD4, gp120 binds to a coreceptor via two principal interactions: (i) the bridging sheet, a four-stranded antiparallel beta sheet that connects the inner and outer domains of gp120, together with the base of the V3 loop, engages the coreceptor N terminus (NT); and (ii) more distal regions of V3 interact with the coreceptor extracellular loops (ECLs) (13, 14, 36-38, 43, 59, 60, 78, 79, 88). Although both the NT and ECL interactions are important for coreceptor binding and entry, their relative contributions vary among different HIV-1 strains (23). For example, V3 interactions with the ECLs, particularly ECL2, serve a dominant role in CXCR4 utilization (7, 21, 50, 63, 72), while R5 viruses exhibit a more variable use of CCR5 domains, with the NT interaction being particularly important (4, 6, 20, 67, 83). Although V3 is the primary determinant of coreceptor preference (34), it is unclear how specificity for CCR5 and/or CXCR4 is determined, and, in particular, it is unknown how X4 tropism is acquired. Several reports have shown that the emergence of X4 tropism correlates with the acquisition of positively charged residues in the V3 stem (17, 29, 87), particularly at positions 11, 24, and 25 (8, 17, 28, 29, 42, 75), raising the possibility that these mutations directly or indirectly mediate interactions with negatively charged residues in the CXCR4 ECLs. However, Env domains outside V3, including V1/V2 (9, 32, 45, 46, 61, 64, 65, 80, 95) and even gp41 (40), can also contribute to coreceptor switching, and it is unclear mechanistically or structurally how X4 tropism is determined.We previously derived a replication-competent variant of the R5X4 HIV-1 clone R3A that contained a markedly truncated V3 loop (47). This Env was generated by introducing a mutation termed ΔV3(9,9), which deleted the distal 15 amino acids of V3. The ΔV3(9,9) mutation selectively ablated X4 tropism but left R5 tropism intact, consistent with the view that an interaction between the distal half of V3 and the ECLs is critical for CXCR4 usage (7, 21, 43, 50, 59, 60, 63, 72). This V3-truncated virus provided a unique opportunity to address whether CXCR4 utilization could be regained on a background in which this critical V3-ECL interaction had been ablated and, if so, by what mechanism. Here, we characterize two novel X4 variants of R3A ΔV3(9,9) derived by adapting this virus to replicate in CXCR4+ CCR5 SupT1 cells. We show that R3A ΔV3(9,9) could indeed reacquire X4 tropism but through two markedly different mechanisms. One X4 variant, designated ΔV3-X4A, acquired changes in the V3 remnant that reestablished an interaction with the CXCR4 ECLs; the other, ΔV3-X4B, acquired changes outside V3 that engendered interactions with the CXCR4 NT. These divergent evolutionary pathways led to profound differences in sensitivity to the CXCR4 antagonist AMD3100, with ΔV3-X4A showing increased sensitivity relative to R3A and with ΔV3-X4B becoming completely resistant. These findings demonstrate the contributions that interactions with distinct coreceptor regions have in mediating tropism and drug sensitivity and illustrate how HIV''s remarkable evolutionary plasticity in adapting to selection pressures can be exploited to better understand its biological potential.  相似文献   

8.
9.
Herpes simplex virus type 1 (HSV-1)-induced cell fusion is mediated by viral glycoproteins and other membrane proteins expressed on infected cell surfaces. Certain mutations in the carboxyl terminus of HSV-1 glycoprotein B (gB) and in the amino terminus of gK cause extensive virus-induced cell fusion. Although gB is known to be a fusogenic glycoprotein, the mechanism by which gK is involved in virus-induced cell fusion remains elusive. To delineate the amino-terminal domains of gK involved in virus-induced cell fusion, the recombinant viruses gKΔ31-47, gKΔ31-68, and gKΔ31-117, expressing gK carrying in-frame deletions spanning the amino terminus of gK immediately after the gK signal sequence (amino acids [aa] 1 to 30), were constructed. Mutant viruses gKΔ31-47 and gKΔ31-117 exhibited a gK-null (ΔgK) phenotype characterized by the formation of very small viral plaques and up to a 2-log reduction in the production of infectious virus in comparison to that for the parental HSV-1(F) wild-type virus. The gKΔ31-68 mutant virus formed substantially larger plaques and produced 1-log-higher titers than the gKΔ31-47 and gKΔ31-117 mutant virions at low multiplicities of infection. Deletion of 28 aa from the carboxyl terminus of gB (gBΔ28syn) caused extensive virus-induced cell fusion. However, the gBΔ28syn mutation was unable to cause virus-induced cell fusion in the presence of the gKΔ31-68 mutation. Transient expression of a peptide composed of the amino-terminal 82 aa of gK (gKa) produced a glycosylated peptide that was efficiently expressed on cell surfaces only after infection with the HSV-1(F), gKΔ31-68, ΔgK, or UL20-null virus. The gKa peptide complemented the gKΔ31-47 and gKΔ31-68 mutant viruses for infectious-virus production and for gKΔ31-68/gBΔ28syn-mediated cell fusion. These data show that the amino terminus of gK modulates gB-mediated virus-induced cell fusion and virion egress.Herpes simplex virus type 1 (HSV-1) specifies at least 11 virally encoded glycoproteins, as well as several nonglycosylated and lipid-anchored membrane-associated proteins, which serve important functions in virion infectivity and virus spread. Although cell-free enveloped virions can efficiently spread viral infection, virions can also spread by causing cell fusion of adjacent cellular membranes. Virus-induced cell fusion, which is caused by viral glycoproteins expressed on infected cell surfaces, enables transmission of virions from one cell to another, avoiding extracellular spaces and exposure of free virions to neutralizing antibodies (reviewed in reference 56). Most mutations that cause extensive virus-induced cell-to-cell fusion (syncytial or syn mutations) have been mapped to at least four regions of the viral genome: the UL20 gene (5, 42, 44); the UL24 gene (37, 58); the UL27 gene, encoding glycoprotein B (gB) (9, 51); and the UL53 gene, coding for gK (7, 15, 35, 53, 54, 57).Increasing evidence suggests that virus-induced cell fusion is mediated by the concerted action of glycoproteins gD, gB, and gH/gL. Recent studies have shown that gD interacts with both gB and gH/gL (1, 2). Binding of gD to its cognate receptors, including Nectin-1, HVEM, and others (12, 29, 48, 59, 60, 62, 63), is thought to trigger conformation changes in gH/gL and gB that cause fusion of the viral envelope with cellular membranes during virus entry and virus-induced cell fusion (32, 34). Transient coexpression of gB, gD, and gH/gL causes cell-to-cell fusion (49, 68). However, this phenomenon does not accurately model viral fusion, because other viral glycoproteins and membrane proteins known to be important for virus-induced cell fusion are not required (6, 14, 31). Specifically, gK and UL20 were shown to be absolutely required for virus-induced cell fusion (21, 46). Moreover, syncytial mutations within gK (7, 15, 35, 53, 54, 57) or UL20 (5, 42, 44) promote extensive virus-induced cell fusion, and viruses lacking gK enter more slowly than wild-type virus into susceptible cells (25). Furthermore, transient coexpression of gK carrying a syncytial mutation with gB, gD, and gH/gL did not enhance cell fusion, while coexpression of the wild-type gK with gB, gD, and gH/gL inhibited cell fusion (3).Glycoproteins gB and gH are highly conserved across all subfamilies of herpesviruses. gB forms a homotrimeric type I integral membrane protein, which is N glycosylated at multiple sites within the polypeptide. An unusual feature of gB is that syncytial mutations that enhance virus-induced cell fusion are located exclusively in the carboxyl terminus of gB, which is predicted to be located intracellularly (51). Single-amino-acid substitutions within two regions of the intracellular cytoplasmic domain of gB were shown to cause syncytium formation and were designated region I (amino acid [aa] positions 816 and 817) and region II (aa positions 853, 854, and 857) (9, 10, 28, 69). Furthermore, deletion of 28 aa from the carboxyl terminus of gB, disrupting the small predicted alpha-helical domain H17b, causes extensive virus-induced cell fusion as well as extensive glycoprotein-mediated cell fusion in the gB, gD, and gH/gL transient-coexpression system (22, 49, 68). The X-ray structure of the ectodomain of gB has been determined and is predicted to assume at least two major conformations, one of which may be necessary for the fusogenic properties of gB. Therefore, perturbation of the carboxyl terminus of gB may alter the conformation of the amino terminus of gB, thus favoring one of the two predicted conformational structures that causes membrane fusion (34).The UL53 (gK) and UL20 genes encode multipass transmembrane proteins of 338 and 222 aa, respectively, which are conserved in all alphaherpesviruses (15, 42, 55). Both proteins have multiple sites where posttranslational modification can occur; however, only gK is posttranslationally modified by N-linked carbohydrate addition (15, 35, 55). The specific membrane topologies of both gK and UL20 protein (UL20p) have been predicted and experimentally confirmed using epitope tags inserted within predicted intracellular and extracellular domains (18, 21, 44). Syncytial mutations in gK map predominantly within extracellular domains of gK and particularly within the amino-terminal portion of gK (domain I) (18), while syncytial mutations of UL20 are located within the amino terminus of UL20p, shown to be located intracellularly (44). A series of recent studies have shown that HSV-1 gK and UL20 functionally and physically interact and that these interactions are necessary for their coordinate intracellular transport and cell surface expression (16, 18, 21, 26, 45). Specifically, direct protein-protein interactions between the amino terminus of HSV-1 UL20 and gK domain III, both of which are localized intracellularly, were recently demonstrated by two-way coimmunoprecipitation experiments (19).According to the most prevalent model for herpesvirus intracellular morphogenesis, capsids initially assemble within the nuclei and acquire a primary envelope by budding into the perinuclear spaces. Subsequently, these virions lose their envelope through fusion with the outer nuclear lamellae. Within the cytoplasm, tegument proteins associate with the viral nucleocapsid and final envelopment occurs by budding of cytoplasmic capsids into specific trans-Golgi network (TGN)-associated membranes (8, 30, 47, 70). Mature virions traffic to cell surfaces, presumably following the cellular secretory pathway (33, 47, 61). In addition to their significant roles in virus-induced cell fusion, gK and UL20 are required for cytoplasmic virion envelopment. Viruses with deletions in either the gK or the UL20 gene are unable to translocate from the cytoplasm to extracellular spaces and accumulated as unenveloped virions in the cytoplasm (5, 15, 20, 21, 26, 35, 36, 38, 44, 55). Current evidence suggests that the functions of gK and UL20 in cytoplasmic virion envelopment and virus-induced cell fusion are carried out by different, genetically separable domains of UL20p. Specifically, UL20 mutations within the amino and carboxyl termini of UL20p allowed cotransport of gK and UL20p to cell surfaces, virus-induced cell fusion, and TGN localization, while effectively inhibiting cytoplasmic virion envelopment (44, 45).In this paper, we demonstrate that the amino terminus of gK expressed as a free peptide of 82 aa (gKa) is transported to infected cell surfaces by viral proteins other than gK or UL20p and facilitates virus-induced cell fusion caused by syncytial mutations in the carboxyl terminus of gB. Thus, functional domains of gK can be genetically separated, as we have shown previously (44, 45), as well as physically separated into different peptide portions that retain functional activities of gK. These results are consistent with the hypothesis that the amino terminus of gK directly or indirectly interacts with and modulates the fusogenic properties of gB.  相似文献   

10.
11.
12.
Imprinted gene expression corresponds to parental allele-specific DNA CpG methylation and chromatin composition. Histone tail covalent modifications have been extensively studied, but it is not known whether modifications in the histone globular domains can also discriminate between the parental alleles. Using multiplex chromatin immunoprecipitation-single nucleotide primer extension (ChIP-SNuPE) assays, we measured the allele-specific enrichment of H3K79 methylation and H4K91 acetylation along the H19/Igf2 imprinted domain. Whereas H3K79me1, H3K79me2, and H4K91ac displayed a paternal-specific enrichment at the paternally expressed Igf2 locus, H3K79me3 was paternally biased at the maternally expressed H19 locus, including the paternally methylated imprinting control region (ICR). We found that these allele-specific differences depended on CTCF binding in the maternal ICR allele. We analyzed an additional 11 differentially methylated regions (DMRs) and found that, in general, H3K79me3 was associated with the CpG-methylated alleles, whereas H3K79me1, H3K79me2, and H4K91ac enrichment was specific to the unmethylated alleles. Our data suggest that allele-specific differences in the globular histone domains may constitute a layer of the “histone code” at imprinted genes.Imprinted genes are defined by the characteristic monoallelic silencing of either the paternally or maternally inherited allele. Most imprinted genes exist in imprinted gene clusters (10), and these clusters are usually associated with one or more differentially methylated regions (DMRs) (27, 65). DNA methylation at DMRs is essential for the allele-specific expression of most imprinted genes (31). Maternal or paternal allele-specific DNA methylation of a subset of DMRs (germ line DMRs) is gamete specific (27, 39). These maternal or paternal methylation differences are established during oogenesis or spermatogenesis, respectively, by the de novo DNA methyltransferases Dnmt3a and Dnmt3b together with Dnmt3L (5, 26, 48). The gamete-specific methylation differences set the stage for the parental allele-specific action of germ line DMRs, some of which have been shown to control the monoallelic expression of the associated genes in the respective domains (11, 34, 36, 53, 66, 71-73, 77). These DMRs are called imprinting control regions (ICRs).Two recurring themes have been reported for ICR action. ICRs can function as DNA methylation-regulated promoters of a noncoding RNA or as methylation-regulated insulators. Recent evidence suggests that both of these mechanisms involve chromatin organization by either the noncoding RNA (45, 50) or the CTCF insulator protein (17, 32) along the respective imprinted domains. The CTCF insulator binds in the unmethylated maternal allele of the H19/Igf2 ICR and blocks the access of the Igf2 promoters to the shared downstream enhancers. CTCF cannot bind in the methylated paternal ICR allele; hence, here the Igf2 promoters have access to the enhancers (4, 18, 24, 25, 62). When CTCF binding is abolished in the ICR of the maternal allele, Igf2 expression becomes biallelic, and H19 expression is missing from both alleles (17, 52, 58, 63). Importantly, CTCF is the single major organizer of the allele-specific chromatin along the H19/Igf2 imprinted domain (17). Significantly, CTCF recruits, at a distance, Polycomb-mediated H3K27me3 repressive marks at the Igf2 promoter and at the Igf2 DMRs (17, 32).A role for chromatin composition is suggested in the parental allele-specific expression of imprinted genes. Repressive histone tail covalent modifications, such as H3K9me2 H3K9me3, H4K20me3, H3K27me3, and the symmetrically methylated H4R3me2 marks, are generally associated with the methylated DMR alleles, while activating histone tail covalent modifications, such as acetylated histone tails and also H3K4me2 and H3K4me3, are characteristic of the unmethylated alleles (7-9, 12-15, 17, 21, 33, 35, 43, 44, 51, 55, 56, 67, 69, 74, 75). Importantly, the maintenance of imprinted gene expression depends on the allele-specific chromatin differences. ICR-dependent H3K9me2 and H3K27me3 enrichment in the paternal allele (67) is required for paternal repression of a set of imprinted genes along the Kcnq1 imprinted domain in the placenta (30). Imprinted Cdkn1c and Cd81 expression depends on H3K27 methyltransferase Ezh2 activity in the extraembryonic ectoderm (64). Similarly, H3K9 methyltransferase Ehmt2 is required for parental allele-specific expression of a number of imprinted genes, including Osbpl5, Cd81, Ascl2, Tfpi2, and Slc22a3 in the placenta (44, 45, 70).There is increasing evidence that covalent modifications, not only in the histone tails but also in the histone globular domains, carry essential information for development and gene regulation. The H3K79 methyltransferase gene is essential for development in Drosophila (60) and in mice (22). H3K79 methylation is required for telomeric heterochromatin silencing in Drosophila (60), Saccharomyces cerevisiae (47, 68), and mice (22). The H4K91 residue regulates nucleosome assembly (76). Whereas mutations at single acetylation sites in the histone tails have only minor consequences, mutation of the H4K91 site in the histone H4 globular domain causes severe defects in silent chromatin formation and DNA repair in yeast (37, 42, 76).Contrary to the abundant information that exists regarding the allele-specific chromatin composition at DMRs of imprinted genes, no information is available about the parental allele-specific marking in the histone globular domains at the DMRs. We hypothesized that chromatin marks in the globular domains of histones also distinguish the parental alleles of germ line DMRs. In order to demonstrate this, we measured the allele-specific enrichment of H3K79me1, H3K79me2, H3K79me3, and H4K91ac at 11 mouse DMRs using quantitative multiplex chromatin immunoprecipitation-single nucleotide primer extension (ChIP-SNuPE) assays. In general, H3K79me3 was associated with the methylated allele at most DMRs, whereas the unmethylated allele showed enrichment for H3K79me1, H3K79me2, and H4K91ac. These results are consistent with the possibility that allele-specific differences in the globular domains of histones contribute to the “histone code” at DMRs.  相似文献   

13.
14.
The glyoxylate and methylcitrate cycles are involved in the metabolism of two- or three-carbon compounds in fungi. To elucidate the role(s) of these pathways in Gibberella zeae, which causes head blight in cereal crops, we focused on the functions of G. zeae orthologs (GzICL1 and GzMCL1) of the genes that encode isocitrate lyase (ICL) and methylisocitrate lyase (MCL), respectively, key enzymes in each cycle. The deletion of GzICL1 (ΔGzICL1) caused defects in growth on acetate and in perithecium (sexual fruiting body) formation but not in virulence on barley and wheat, indicating that GzICL1 acts as the ICL of the glyoxylate cycle and is essential for self-fertility in G. zeae. In contrast, the ΔGzMCL1 strains failed to grow on propionate but exhibited no major changes in other traits, suggesting that GzMCL1 is required for the methylcitrate cycle in G. zeae. Interestingly, double deletion of both GzICL1 and GzMCL1 caused significantly reduced virulence on host plants, indicating that both GzICL1 and GzMCL1 have redundant functions for plant infection in G. zeae. Thus, both GzICL1 and GzMCL1 may play important roles in determining major mycological and pathological traits of G. zeae by participating in different metabolic pathways for the use of fatty acids.During the infection process, pathogenic fungi usually encounter nutrient deprivation in the host before gaining access to sufficient nutrients for successful colonization of the living tissue. To cope with a nutrient-limited environment, fungal pathogens seem to rely mostly on fatty acid metabolism for both energy supply and biosynthesis of essential molecules (29). The ability of fungi to use fatty acids as a carbon source for growth is based on the glyoxylate cycle. Fungal pathogens have been proposed to employ the glyoxylate bypass for the use of acetyl coenzyme A (CoA) units produced by the β-oxidation of even-chain-length fatty acids, probably available from host cell membranes or the lipid reservoir inside the fungal spore (7, 12, 20, 27, 28, 41, 44, 46). Recent studies suggest that the glyoxylate pathway plays an important role in fungal virulence toward both plant and animal hosts (12, 20, 27, 44, 46). The key enzymes of the glyoxylate pathway, such as isocitrate lyase (ICL), which catalyzes the cleavage of isocitrate to glyoxylate and succinate, and malate synthase, which mediates the condensation of acetyl-CoA and glyoxylate into malate, are strongly induced within the host (16, 27, 41, 44). Moreover, disruption of genes encoding either of these enzymes causes severely reduced virulence of fungal phytopathogens, including Leptosphaeria maculans (20), Magnaporthe grisea (46), Stagonospora nodorum (44), and Colletotrichum lagenarium (2), and the animal pathogen Candida albicans (27). In contrast, these glyoxylate cycle enzymes have been known to be dispensable in invasive aspergillosis caused by Aspergillus fumigatus (38, 43).During fatty acid and amino acid catabolism by fungi, propionyl-CoA can be generated along with acetyl-CoA, particularly from the breakdown of odd-chain-length fatty acids or of the amino acids valine, isoleucine, and methionine (14). Therefore, fungal pathogens may need to use or remove propionyl-CoA during the infection process because it is toxic to fungi. In fungi, propionyl-CoA is metabolized via the methylcitrate cycle, in which propionyl-CoA is oxidized to pyruvate in four enzymatic steps (4, 5, 6, 19, 30, 31, 40, 49, 50). Recently, the importance of the methylcitrate cycle in fungal virulence was demonstrated in A. fumigatus: a mutant defective in methylcitrate synthase, the first enzyme of this cycle, displayed attenuated virulence in mice and insects (19, 31). However, the role of methylisocitrate lyase (MCL), which catalyzes the last reaction in the methylcitrate cycle (i.e., the cleavage of methylisocitrate into pyruvate and succinate) in fungal virulence, has not been determined, although deletion of the MCL gene inhibits hyphal growth and conidiation in Aspergillus nidulans (4). The protein sequences of several fungal MCLs show high similarity to fungal ICLs of the glyoxylate cycle (4, 30). In the pathogenic bacterium Mycobacterium tuberculosis, the methylcitrate cycle, only when working together with the glyoxylate cycle, is involved in virulence as well as fatty acid metabolism and intracellular growth (34, 35).Here, we focused on the roles of these two cycles during disease development caused by the devastating cereal pathogen Gibberella zeae (anamorph: Fusarium graminearum). G. zeae is a ubiquitously distributed ascomycete fungus that causes major disease in cereal crops such as corn, wheat, barley, and rice (33). Severe epidemics of these diseases result in serious economic consequences due to yield losses and contamination by fungal mycotoxins (32, 33). Wind-disseminated sexual spores (ascospores), which are produced in perithecia formed on plant debris, can infect plant spikes during anthesis (13, 39, 45). Detailed studies of the G. zeae infection process on wheat and barley heads have shown that fungal hyphae on the inner surfaces of the spike penetrate epicarp cells through pits or pores and grow into the caryopses through the pericarp (21). Thus, the glyoxylate cycle, either alone or in conjunction with the methylcitrate cycle, is likely employed by G. zeae during the infection process, as in other fungus-plant interactions (20, 46). G. zeae genome searches have identified orthologs of fungal ICL and MCL genes, designated GzICL1 and GzMCL1, respectively. Here, we performed functional analyses of these genes to provide new insight into their importance in lipid metabolism during the G. zeae infection process in host plants.  相似文献   

15.
16.
17.
18.
19.
20.
An attenuated derivative of simian immunodeficiency virus strain 239 deleted of V1-V2 sequences in the envelope gene (SIV239ΔV1-V2) was used for vaccine/challenge experiments in rhesus monkeys. Peak levels of viral RNA in plasma of 104 to 106.5 copies/ml in the weeks immediately following inoculation of SIV239ΔV1-V2 were 10- to 1,000-fold lower than those observed with parental SIV239 (∼107.3 copies/ml). Viral loads consistently remained below 200 copies/ml after 8 weeks of infection by the attenuated SIV239ΔV1-V2 strain. Viral localization experiments revealed large numbers of infected cells within organized lymphoid nodules of the colonic gut-associated lymphoid tissue at 14 days; double-labeling experiments indicated that 93.5% of the virally infected cells at this site were positive for the macrophage marker CD68. Cellular and humoral immune responses measured principally by gamma interferon enzyme-linked immunospot and neutralization assays were variable in the five vaccinated monkeys. One monkey had responses in these assays comparable to or only slightly less than those observed in monkeys infected with parental, wild-type SIV239. Four of the vaccinated monkeys, however, had low, marginal, or undetectable responses in these same assays. These five vaccinated monkeys and three naïve control monkeys were subsequently challenged intravenously with wild-type SIV239. Three of the five vaccinated monkeys, including the one with strong anti-SIV immune responses, were strongly protected against the challenge on the basis of viral load measurements. Surprisingly, two of the vaccinated monkeys were strongly protected against SIV239 challenge despite the presence of cellular anti-SIV responses of low-frequency and low-titer anti-SIV antibody responses. These results indicate that high-titer anti-SIV antibody responses and high-frequency anti-SIV cellular immune responses measurable by standard assays from the peripheral blood are not needed to achieve strong vaccine protection, even against a difficult, neutralization-resistant strain such as SIV239.The characteristics of human immunodeficiency virus type 1 (HIV-1) infection suggest major difficulty for the development of a preventive vaccine (19, 23). Pessimism regarding the prospects for a vaccine is derived at least in part from the ability of HIV-1 to continually replicate in the face of apparently strong host immune responses, resistance to antibody-mediated neutralization, and the extensive sequence diversity in field strains of the virus. Lack of knowledge regarding the key components of a protective immune response also remains a major scientific obstacle. Vaccine/challenge experiments with macaque monkeys have been used to evaluate the properties and relative effectiveness of different vaccine approaches and to gauge the formidable nature of these difficulties.One lesson that has been learned from vaccine/challenge experiments with macaque monkeys is the importance of challenge strain on outcome. Vaccinated monkeys that have been challenged with strains of simian immunodeficiency virus (SIV) with an HIV-1 envelope (SHIV) have almost invariably exhibited strong, long-term protection against disease, irrespective of the nature of the vaccine. Even peptide immunogens have protected against SHIV-induced disease (6, 12, 38). Vaccine approaches that have protected against SHIV challenge include DNA (5, 13), recombinant poxvirus (4), recombinant adenovirus (57), other viral recombinants (18, 55), prime and boost protocols (3, 53, 65), and purified protein (10, 64). Vaccine protection against pathogenic SIV strains such as SIV239, SIV251, and SIV-E660 has been much more difficult to achieve (2, 11, 27, 63). The identical replication-defective gag-recombinant adenovirus that provided strong protection against SHIV challenge (57) provided little or no protection against SIV239 challenge (11). Disappointing levels of protection against SIV have often been observed in the face of apparently robust vaccine-induced immune responses (see, for example, Vogel et al. [63] and Casimiro et al. [11]). Some partial vaccine protections against these SIV strains have been achieved by recombinant poxvirus (7, 50), replication-competent recombinant adenovirus (51), replication-defective adenovirus (66), recombinant poliovirus (15), recombinant Venezuelan equine encephalitis virus (18), and recombinant Sendai virus (44).Differences between the biological properties of the SIV strains and those of the SHIV strains used for the above-mentioned studies provide clues as to what may be responsible for the differences in outcome. These SIV strains are difficult to neutralize (26, 34), use CCR5 as a coreceptor for entry into cells (21, 52), and induce a chronic, progressive disease course (17), and this course is independent of the infectious dose (17). The SHIV strains used for the above-mentioned studies are easier to neutralize, use CXCR4 for entry, and induce an acute decline in CD4 counts, and the disease course is dose dependent (29, 30, 48, 54). These SIV strains, like HIV-1 in humans, exhibit a marked preference for CD4+ CCR5+ memory cells, in contrast to the acutely pathogenic SHIV strains which principally target naïve cells (48).Live, attenuated strains of SIV have provided the strongest vaccine protection by far against SIV challenge. Although clinical use of a live, attenuated HIV vaccine is not being considered, understanding the basis of the strong protection afforded by live, attenuated SIV strains remains an important research objective for the insights that can be provided. Most of the attenuated SIV strains that have been used lack a functional nef gene (16, 31, 58, 67). Shacklett et al. (56) used an attenuated SIV strain with modifications in the gp41 transmembrane protein for protection. Here, we describe strong vaccine protection by a replication-competent SIV strain lacking 100 amino acids from the essential gp120 envelope protein in the absence of overtly robust immune responses.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号